Oxidative Stress - American Chemical Society

B-doped diamond electrodes (38) but lower cost methods based on resonance elastic light .... 0, (2) 0.4, (3) 0.79, (4) 1.6, (5) 2.3, (6) 3.1, (7) 3.85...
1 downloads 0 Views 2MB Size
Chapter 14

Redox Activity of Oxidative Stress-Damping Endogenous Thiol Biomolecules Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Agata Chalupa and Maria Hepel* Department of Chemistry, State University of New York at Potsdam, Potsdam, New York 13676 *E-mail: [email protected]

The internal oxidative-stress control and prevention system in living organisms is based on homeostasis of redox potential. The main effector of this system is the redox couple: glutathione (GSH) and its disulfide (GSSG), with the support of NADPH and GSH reductase. GSH in this redox couple is a ubiquitous reducing agent responsible for neutralization of harmful radicals and reactive oxygen species. However, the processes of GSH oxidation and GSSG reduction on electrodes are strongly hindered. In this Chapter, we present results of our investigations of redox reactivity of endogenous thiols: GSH, cysteine (Cys), and homocysteine (Hcys) on electrocatalytic cobalt phthalocyanine (CoPc) monolayer film on glassy carbon electrode. We show that a strong competition between GSH, Cys, and Hcys exists similar to that observed in thiol-capped Au nanoparticle assembly processes. Detailed investigations indicate that the electroanalytical determination of thiols in multicomponent solutions must take into account the dependence of sensitivity on competitive charge-transfer complex formation.

Introduction The cellular oxidative stress originates from various internal and external sources. Any process that acts to increase the production of reactive oxygen species (ROS) (1, 2), including H2O2, O2•-, and HO•, and biomolecular radicals, is causing the oxidative stress. The oxidative stress has a profound impact on © 2015 American Chemical Society In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

life processes (3), including the metabolism, energy production, communication pathways, cell cycles, anti-pathogenic defences, and disease control. While the oxidative stress performs an outstanding life-supporting and cellular defense duties, it may also act toward organ damage and disease progression. In the latter category, oxidative stress-generating ROS radicals may participate in carcinogenesis and, for instance, by inducing inflamamtion of heart muscle leading to cardiomyopathy and coronary heart disease (4, 5). The excess of ROS in the brain is damaging the central nervous system leading to such illnesses as the Parkinson’s (6–8), Down syndrome (9, 10), and autism (11, 12). It has been found that the oxidative stress may cause a serious damage to DNA (13, 14), proteins, and lipids (3). It contributes to many debilitating or fatal diseases, such as diabetes (15, 16), Alzheimer’s (17–19), cardiovascular (4, 5, 20, 21), acute renal failure (22, 23), cancer (24, 25), and others, and plays an important role in accelerating the aging process (3). The devastating effects of ROS overproduction can be counterbalanced by anti-oxidants, the compounds scavenging biomolecule radicals and ROS. There are many groups of exogeneous anti-oxidants that are consumed with foods, including flavonoids, polyphenols, phytochemicals, etc. The internal defence system of most living organisms against radical attacks is based on a small biothiol molecule, glutathione (GSH) (26) which is a strong reducing agent. GSH undergoes oxidation to glutathione disulfide (GSSG) giving up an electron to a radical, and thereby neutralizing it. The concentration ratio of GSH to GSSG determines the anti-oxidizing power of the couple (27, 28). The depleted GSH can be replenished by the reaction of GSSG with NADPH catalyzed by GSH reductase participating in the redox level homeostasis (29, 30). The sulfhydryl groups of proteins and other biothiol molecules, such as cysteine (Cys), can also participate in maintaining the low redox potential and thus prevent oxidative stress. Analytical determinations of biothiols can be carried out by different techniques. Excellent reviews on this subject have been published (31, 32). The determination of GSH and Cys, as well as homocysteine (Hcys), can be performed using HPLC with fluorescence detection (33, 34), liquid chromatography-mass spectroscopy (LC-MS) (35, 36), gas chromatography-mass spectroscopy (GC-MS) (35, 37) or B-doped diamond electrodes (38) but lower cost methods based on resonance elastic light scattering (39–43), UV-Vis absorption (44), fluorescence (45–48), electrochemical techniques (49–52) and on sensors (53–61) are also available. Lee et al. (62) have developed an electrochemical method for detection of GSH and homocysteine by expoiting differences in the rates of their reactions with an in situ formed oxidized catechol. The 1,4-Michael addition reaction leads to the formation of an adduct exhibiting a new voltammetric peak which was utilized for sensitive thiol determination. Unfortunately, many electroanalytical methods require extensive pretreatment procedures (63–75) to attain sufficient sensitivity. Often, different reagents have been used to either derivatize biothiols or modify their electrochemical activity to achieve better selectivity (76–81). Recently, a method for the determination of total concentration of antioxidants: GSH, cysteine, homocysteine, and ascorbic acid, has been developed by Compton et al. (82). Distinguishing between GSH, Cys, and Hcys has been attempted (39, 40, 43, 83–85) but still remains a challenge, especially due to the structural similarity 330 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

of Cys and Hcys. The Au and Ag colloids have been widely used to study the interactions of biothiols with plasmonic nanoparticles (39, 40, 86, 87) and to evaluate the ligand exchange processes in nanoparticle-shell self-assembling monolayers using colorimetric or RELS techniques. The redox reactivity of GSH/GSSG couple in homogeneous biological media is fast due to the enzymatic lowering of activation energy barrier of electron transfer processes. However, on solid electrodes, both the oxidation of GSH and the reduction of GSSG are strongly hindered (39). In this Chapter, we present results of investigations of redox reactivity of glutathione, cysteine, and homocysteine on an electrocatalytic cobalt phthalocyanine (CoPc) monolayer film on a glassy carbon electrode (GCE). The studies of GSH oxidation of CoPc have earlier been carried out by Zagal and coworkers (53) on CoPc-coated graphite electrodes and polypyrrole-embedded CoPc sensors (55), and recently in our group on CoPc-modified glassy carbon and graphene electrodes. We show that a strong competition between GSH, Cys, and Hcys exists similar to that observed in thiol-capped Au nanoparticle assembly processes (39, 88) which prevents a straightforward analysis of multicomponent mixtures of the thiols. Therefore, we have performed detailed investigations to develop a framework for the multicomponent analysis taking into account active matrix effects. The remarkable activity of CoPC acting as an artificial enzyme for oxidation of biothiols and their disulfide reduction may be utilized in future for developing sensitive, fast, and inexpensive biomimetic sensors for screening biomarkers of oxidative stress in doctors office and field health centers.

Experimental Chemicals All chemicals used for investigations were of analytical grade purity. Reduced L-glutathione (GSH), DL-homocysteine (Hcys), L-cysteine (Cys), N,N-Dimethylformamide (DMF), boric acid, acetic acid, and cobalt(II) phthalocyanine were purchased from Sigma-Aldrich Chemical Company (St. Louis, MO, U.S.A.). Sodium phosphate dibasic heptahydrate (Na2HPO4•7 H2O), phosphoric acid and sodium phosphate monobasic dihydrate (NaH2PO4•2 H2O) were obtained from J.T. Baker Chemical Co. Solutions were prepared using Millipore (Billerica, MA, U.S.A.) Milli-Q deionized water (conductivity σ = 55 nS/cm). They were deoxygenated by bubbling with purified nitrogen.

Instrumentation Cyclic voltammetric (CV) measurements were performed with Elchema potentiostat/galvanostat Model PS-205 (Potsdam, NY, USA) with a threeelectrode configuration. Potentials were measured versus the double-junction Ag/AgCl reference electrode, obtained from Elchema, with a 3 M KNO3 external filling solution. A Pt wire was used as the counter electrode. The glassy carbon 331 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

electrode (GCE) with an area of 7.1 mm2, obtained from Elchema (Potsdam, NY, U.S.A.) was used as the working electrode. The program waveform was supplied and data acquisition performed by Voltscan 5.0 interactive software from Elchema with 16-bit precision. The measurements were carried out in 50 mM phosphate buffer solutions, pH 7.4, unless otherwise stated. The surface of GCE electrode was polished with 0.05 µm alumina slurry (Coating Service Department, Indianapolis, U.S.A.) on a flat pad and rinsed repeatedly with water to remove any alumina residue. The potential was scanned between -1.1 V and +0.4 V, unless otherwise noted. Subsequent measurements were carried out by polishing the surface of GCE and depositing a fresh electrocatalyst coating.

Molecular Dynamics Simulation and Quantum Mechanical Calculations Quantum mechanical calculations of electronic structures for CoPc molecule and its interactions with GSH, cysteine and homocysteine were performed using modified Hartree-Fock methods with 6-31G* basis set and pseudopotentials, semi-empirical PM3 method, and density functional theory (DFT) with B3LYP functional and correlated methods (89, 90). The molecular dynamics simulations and quantum mechanical calculations were carried out using procedures embedded in Wavefunction (Irvine, CA, U.S.A.) Spartan 6. The electron density and local density of states are expressed in atomic units, au-3, where 1 au = 0.52916 Å and 1 au-3 = 6.7491 Å-3.

Results and Discussion Cobalt-Phthalocyanine-Coated Catalyst Sensor The structures of main compounds studied are presented in Scheme 1, including cobalt phthalocyanine (CoPc), glutathione (GSH), homocysteine (Hcys), and cysteine (Cys). Note that Co cation in CoPc complex is formally on the +1 or +2 oxidation state. Upon adsorption or immobilization of CoPc on the electrode surface, it can be oxidized/reduced by appropriately adjusting the electrode potential. Testing different catalysts for electrooxidation of GSH, Cys, and Hcys, as well as the GSSG reduction, has been performed using derivatives of coumarin, fluorone black, and cobalt phthalocyanine (CoPc) as the electron mediators. Here we present the results obtained with cobalt CoPc electrocatalyst (Scheme 1a and 2). CoPc has unique electron mediation properties and it works specifically with thiols and disulphides. In tests with this catalyst, several electrode materials and different sensor compositions have been prepared and their electroactivity was evaluated. We have found that, the substrates tested, including Au, Pt, and various forms of carbon, do not show any electrocatalytic activity toward thiol oxidation when tested alone. However, when exposed to CoPc, they gain catalytic activity which is then clearly due to the adsorbed CoPc molecules (53, 91, 92). For instance, carbon electrodes coated with CoPc offer much better catalytic 332 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

activity than bare electrodes. Among different carbons, we have examined the following: glassy carbon (GCE), ordinary graphite (OGE), pyrolytic carbon (PCE), carbon nanotube (CNT) printed electrodes, covalently bound CNT (cbCNT), highly oriented pyrolytic graphite (HOPG), and graphene nanosheet substrates (GNS). Here, we present the results for GCE coated with CoPc. From these measurements, best potentials have been selected for each component to enable differentiation. Different levels of catalytic activity come from CoPC and its specific interactions with the carbon substrates.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Characterization of GCE/CoPC The CoPC electrocatalyst layer was deposited by casting from a 50 mM CoPc solution in DMF. The adsorption was carried out for 30 min and after that time the drop of casting solution was dislodged and the sensor was washed several times with DMF and distilled water followed by drying in a stream of nitrogen. In Figure 1, cyclic voltammetry characteristics for a GCE/CoPc electrode is presented in the potential range from -1.15 to +0.4 V vs. Ag/AgCl reference. The electrode capacitance, determined at E = -0.5 V, is C = 209.5 µF/cm2. The high value of differential capacitance is due to the extended real surface area and the catalyst film.

Scheme 1. Structure of the compounds used in the investigations: a) Cobalt phthalocyanine (CoPc); b) glutathione (GSH); c) homocysteine (Hcys), and (d) cysteine. 333 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Scheme 2. Electronic density surface with electrostatic potential map of a CoPc molecule; σ = 0.08, potential color coded from high (blue) to low (red). (see color insert)

Figure 1. (a) Cyclic voltammogram of the cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc); ν = 100 mV/s, electrolyte: 50 mM phosphate buffer, pH 7.43; (b) Dependence of iE=0 vs. ν. The higher slope of the anodic curve (upper line) is due to the contribution of the oxidation current of CoPc.

Electrocatalytic Activity of GCE/CoPc toward GSH Oxidation The catalytic properties of GCE/CoPc electrode are illustrated in Figure 2a. Voltammetric curve 1 was obtained in 3.85 mM GSH solution on bare 334 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

GCE electrode and curve 2 on GCE/CoPc electrode. It is seen that there is no electrocatalytic activity of the GCE alone, but a pronounced catalytic GSH-oxidation current peak is observed on a GCE/CoPc electrode. On the return scan, a cathodic current peak at E = -0.96 V is observed. It is due to the reduction of GSSG formed during the forward scan at potentials E > -0.2 V. The background curve obtained on a bare GCE in buffer solution is presented in curve 3. There is no extensive background shift on adding GSH to the solution. This is important since the background stability contributes to the measurement accuracy at low analyte concentration.

Figure 2. (a) Cyclic voltammograms obtained in 3.85 mM GSH solution for: (1) a bare glassy carbon electrode (GCE) and (2) a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc). Curve (3) was obtained for a bare GCE electrode in buffer solution. (b) Cyclic voltammograms obtained for a GCE/CoPc electrode in solutions with concentrations of GSH, CGSH [mM]: (1) 0, (2) 0.79, (3) 1.6, (4) 2.3, (5) 3.1, (6) 3.85. (c) Dependence of anodic current peak for GSH oxidation (Epa = +0.115 V) on CGSH. (d) Dependence of cathodic peak current for GSSG reduction (Epc = -0.97 V) on CGSH. Other conditions: ν = 100 mV/s; 50 mM phosphate buffer, pH 7.43. (see color insert) 335 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

To confirm that the observed anodic and cathodic currents are due to the oxidation of GSH and reduction of GSSG, respectively, detailed GSH-conecntration studies have been performed. The obtained results are presented in Figures 2-3. A collection of CV curves obtained for a GCE/CoPc electrode in solutions with GSH concentration ranging from 0 to 3.85 mM are presented in Figure 2b. The dependence of anodic current peak at Epa = +0.115 V for GSH oxidation on CGSH is presented in Figure 2c and the dependence of cathodic peak current at Epc = -0.97 V for GSSG reduction on CGSH is presented in Figure 2d.

Figure 3. Cyclic voltammograms obtained for GSH solutions for a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc), CGSH [mM]: (1) 0, (2) 0.4, (3) 0.79, (4) 1.6, (5) 2.3, (6) 3.1, (7) 3.85, (8) 4.5, (9) 6.0, (10) 7.7, (11) 9.4, (12) 11, (13) 12.6; 50 mM phosphate buffer, pH 7.43; ν = 100 mV/s. INSET: Dependence of ip vs. CGSH.

The calibration plots in Figure 2 are linear. However, at higher GSH concentrations (CGSH > 4 mM), a nonlinear behavior is clearly observed. It is illustrated in Figure 3. The dependence of peak current for GSH oxidation and GSG reduction on square root of the scan rate is presented in Figure 4. The plots are linear which confirms that the processes in question involve solution species: GSH and GSSG. 336 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 4. (a) Cyclic voltammograms for 3.85 mM GSH in 50 mM phosphate buffer, pH 7.43 on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc), recorded at ν [mV/s]: (1) 25, (2) 50, (3) 100, (4) 200, (5) 300, (6) 400, (7) 500; (b) dependence of ip vs. sqrt(ν). (see color insert) Note that the redox potential of GSSG/GSH couple is given by:

where the relative standard potential at pH = 7 is (93, 94):

where n = 2, R is the gas constant, T is the absolute temperature, and the Nernstian slope 2.302RT/F = 0.05916 V at 25 0C. This shows that even with the electrocatalyst film on GCE, the oxidation process of GSH is still irreversible. Electrocatalytic Activity of GCE/CoPc toward Hcys Oxidation The GCE/CoPc electrodes have also been tested for the response to homocysteine and cysteine. In Figure 5, the electrocatalytic properties of GCE and GCE/CoPc toward the oxidation of Hcys are compared. Voltammetric curve 1 was obtained in 3.85 mM Hcys solution on bare GCE electrode and curve 2 on GCE/CoPc electrode. It is seen that there is no electrocatalytic activity of the GCE alone, but pronounced catalytic Hcys-oxidation current peaks are observed on a GCE/CoPc electrode. In contrast to GSH, two anodic peaks for the oxidation of Hcys are observed. They likely correspond to the formation of a radical Hcys• intermediate in the first step, followed by the formation of a disulphide in the second step. On the return scan, a cathodic current peak at E = -0.96 V is observed. It is due to the reduction of homocystine (a disulphide) formed during the forward scan at potentials E > +0.18 V. The background curve obtained on 337 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

a bare GCE in buffer solution is presented in curve 3. There is no extensive background shift on adding Hcys to the solution.

Figure 5. Cyclic voltammograms for homocysteine obtained on: (1) bare glassy carbon electrode (GCE) and (2) cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc); CHcys, 3.85 mM. Curve (3) represents the background curve for a bare GCE in phosphate buffer. ν = 100 mV/s, electrolyte: 50 mM phosphate buffer, pH 7.43. To confirm that the observed anodic and cathodic currents are due to the oxidation of Hcys and reduction of homocystine, respectively, detailed Hcys-concentration studies have been performed. The obtained results are presented in Figure 6. A collection of CV curves obtained for a GCE/CoPc electrode in solutions with Hcys concentration ranging from 0 to 3.85 mM are presented in Figure 6. The linear dependence of anodic current peak at Epa = +0.115 V for the first step of Hcys oxidation on CHcys, presented in Figure 6b (upper line), indicates that the reaction is first order with respect to Hcys. The dependence of cathodic peak current at Epc = -0.97 V for homocystine reduction on CHcys, presented in Figure 6b (lower line) is also linear. The scan dependence of anodic and cathodic peak currents has been analyzed in the range of scan rates from 25 to 500 mV/s, as illustrated in Figure 7. Both peak currents depend linearly on square root of scan rate. This confirms that the reactants originate from the solution phase rather than from a film on the electrode. This means that homocystine formed during the anodic oxidation at potentials E > 0.3 V remains in the solution phase in the vicinity of the electrode surface and is available for backward reduction during the cathodic going scan, at potentials E < -0.8 V. The separation between the anodic and cathodic waves is large, ΔEp > 0.9 V (taking into account the first anodic step), indicating on a high irreversibility of the oxidation/reduction processes of Hcys. 338 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 6. (a) Cyclic voltammograms for Hcys in 50 mM phosphate buffer, pH 7.43, recorded on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc), for CHcys [mM]: (1) 0, (2) 0.79, (3) 1.6, (4) 2.3, (5) 3.1, (6) 3.85. (b) Dependence of ip vs. CHcys; ν = 100 mV/s. (see color insert)

Figure 7. (a) Cyclic voltammograms for a 3.85 mM Hcys in 50 mM phosphate buffer, pH 7.43, obtained on a cobalt phthalocyanine modified-glassy carbon electrode (GCE/CoPc) with ν [mV/s]: (1) 25, (2) 50, (3) 100, (4) 200, (5) 300, (6) 400, (7) 500; (b) Dependence of ip vs. square root of ν. The homocysteine voltammetric fingerprint is different than that of GSH, although the reactivity of both compounds is similar. In Figure 8, two families of voltammograms for an increasing anodic reversal potential Era are presented. It is seen that the cathodic peak grows with increasing Era. The growth continues when Era increases in the potential range of the first Hcys oxidation peak and then remains largely invariant with Era when Era increases in the area of the second Hcys oxidation peak. This is likely due to the fact that the Hcys oxidation product, formed in the first stage (i.e. in the potential area of the first oxidation peak), adsorbs on the electrode surface while further oxidation leads to the product that is solution-soluble. These characteristics are important for the analysis of GSSG based on measurements of the rate of its reduction at E < 0.9 V vs Ag/AgCl. 339 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 8. (a) Cyclic voltammograms of Hcys recorded on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) for increasing anodic reversal potential Era [mV]: (1) -200, (2) 0, (3) 100, (4) 200, (5) 300, (6) 400, (7) 500, (8) 550, (9) 600; ν = 100 mV/s, CHcys = 400 µM in 50 mM phosphate buffer, pH 7.43; (b) same, but with Era changing only in the range of the second Hcys oxidation peak. (see color insert)

Electrocatalytic Activity of GCE/CoPc toward Cys Oxidation The GCE/CoPc electrodes have also been tested for the response to cysteine. In Figure 9, the electrocatalytic properties of GCE and GCE/CoPc toward the oxidation of Cys are compared.

Figure 9. Cyclic voltammograms for cysteine (Cys)obtained on: (1) bare glassy carbon electrode (GCE) and (2) cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc); CCys = 790 µM. Curve (3) represents the background curve for bare GCE electrode in phosphate buffer. ν = 100 mV/s, electrolyte: 50 mM phosphate buffer, pH 7.43. 340 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Voltammetric curve 1 was obtained in 3.85 mM Cys solution on bare GCE electrode and curve 2 on GCE/CoPc electrode. It is seen that there is no electrocatalytic activity of the GCE alone, but a pronounced catalytic Cys-oxidation current peak at E = +0.04 V vs. Ag/AgCl is observed on a GCE/CoPc electrode. In contrast to Hcys, only one anodic peak for the oxidation of Cys is observed, despite of the structural and functional similarity of Hcys and Cys (the only difference being one more CH2 group in the Hcys carbon chain). The single Cys oxidation peak means that the disulphide (cystine) is formed in two consecutive overlapping steps. On the return scan, a cathodic current peak at E = -0.97 V is observed. It is due to the reduction of cystine formed during the forward scan at potentials E > -0.2 V. The background curve obtained on a bare GCE in buffer solution is presented in curve 3. There is no extensive background shift on adding Cys to the solution. To confirm that the observed anodic and cathodic currents are due to the oxidation of Cys and reduction of cystine, respectively, detailed Cys-concentration studies have been performed. The obtained results are presented in Figure 10. A collection of CV curves obtained for a GCE/CoPc electrode in solutions with Cys concentration ranging from 0 to 3.85 mM are presented in Figure 10. The linear dependence of anodic current peak at Epa = +0.04 V for the Cys oxidation on CCys, presented in Figure 10b (upper line), indicates that the reaction is first order with respect to Cys. The dependence of cathodic peak current at Epc = -0.97 V for cystine reduction on CCys, presented in Figure 10b (lower line) is also linear.

Figure 10. (a) Cyclic voltammograms, recorded on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc), for Cys solutions with CCys [mM]: (1) 0, (2) 0.79, (3) 1.6, (4) 2.3, (5) 3.1, (6) 3.85; ν = 100 mV/s; 50 mM phosphate buffer, pH 7.43; (b) dependence of ip vs. CCys. (see color insert)

The scan dependence of anodic and cathodic peak currents has been analyzed in the range of scan rates from 25 to 500 mV/s, as illustrated in Figure 11. Both dependencies are linear with square root of scan rate, confirming that reactants from the solution are involved. 341 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 11. (a) Cyclic voltammograms for 3.85 mM Cys in 50 mM phosphate buffer, pH 7.43, on a cobalt phthalocyanine modified glassy carbon electrode (GCE/CoPc), recorded for ν [mV/s]: (1) 25, (2) 50, (3) 100, (4) 200, (5) 300, (6) 400, (7) 500.

Selection of Characteristic Potentials for Differentiation of GSH, Hcys and Cys Whereas GSH, Hcys, and Cys exhibit similar oxidation and reduction processes, their voltammetric characteristics differ in shape and the oxidation and reduction potentials. These differences can be exploited for differentiation between these thiols. Also, the analysis of mixed solutions can be readily performed provided that no appreciable interference between the thiols exists which, however, may not be the case due to the strong competition between thiols. The measurements described below have been carried out to address these issues. In Figure 12, cyclic voltammograms for Hcys, Cys, and GSH have been compared. They were recorded on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) in 50 mM phosphate buffer, pH 7.43, for 790 µM thiols, at a scan rate of 100 mV/s. It is seen that the anodic oxidation current onset is the lowest for Hcys, followed by Cys and GSH. At the same concentration level, Hcys exhibits two anodic peaks at Epa,1 = +0.015 V and Epa,2 = +0.340 V, Cys shows a single anodic peak at Epa = +0.040 V, and GSH a wave with Epa = +0.080 V. The Hcys valley observed in the potential range between the two anodic peaks shows a current minimum at Emin = +0.120 mV. By selecting the four peak potentials and Emin, the current-voltage profile can be analyzed. There are however, subtle problems with matrix effects which have to be addressed. Because of these effects, the standard addition method has to be applied in the analysis. In the following Figures, some of the matrix effects and interdependencies are presented. 342 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 12. Cyclic voltammograms recorded on a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) in 50 mM phosphate buffer, pH 7.43, for: (1) buffer only, (2-4) after addition of the 0.79 mM analyte: (2) Hcys, (3) Cys, and (4) GSH; curve (5) shows the background CV for a bare GCE electrode without catalyst; ν = 100 mV/s. The competition between GSH and Hcys is illustrated in Figure 13. It is seen that the anodic peak currents for oxidation of GSH and Hcys depend on the order in which the reagents (GSH and Hcys) are added. The concentration of Hcys is selected much lower than that of GSH so that the contribution of Hcys to the anodic current is negligible, as shown in curve 1. The experiments of Figure 13a have been performed for a high ratio of GSH to Hcys, CGSH/CHcys = 385. GSH alone shows a high peak current (curve 2). When GSH and Hcys are added to the buffer solution in a different sequence, the GSH peak current is seen to change markedly. Curve 3 shows that if Hcys, despite of its low concentration, is added before GSH, then the GSH peak is 30% lower than that for GSH alone. When GSH and Hcys are added together at the same time (i.e., when both reagents are present in the solution when the electrode is first immersed), the decrease of the GSH peak is smaller. There is clearly a strong competition between GSH and Hcys for the electrocatalytic redox centers in CoPc. These effects have not been described in the literature and this is the first investigation uncovering this competition. While the competitive adsorption of electroactive species and ligands is well known, here we deal rather with a slow-complexation competition since there is virtually no adsorption of reagents on a GCE surface. Similar experiments have been performed for a lower ratio of GSH to Hcys, CGSH/CHcys = 4. The obtained results are presented in Figure 13b, for CGSH = 1.6 mM and CHcys = 400 µM. Curve 3 shows a GSH oxidation current in absence of Hcys and curve 4 shows the oxidation wave when GSH was added first, followed by the addition of Hcys. A higher anodic current is observed due to the additive effect of the oxidation of GSH and Hcys. 343 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 13. (a) Cyclic voltammograms for a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) obtained after addition of: (1) 10 µM Hcys, (2) 3.85 mM GSH, (3) 3.85 mM GSH to a buffer containing 10 µM Hcys, (4) 10 µM Hcys + 3.85 mM GSH added at the same time. (b) Cyclic voltammograms for GCE and GCE/CoPc electrodes: (1) GCE in a buffer solution, (2) GCE/CoPc in a buffer solution, (3) GCE/CoPc in 1.6 mM GSH, (4) GCE/CoPc in 1.6 mM GSH after addition of 400 µM Hcys. Other conditions: 50 mM phosphate buffer solution, pH 7.43, ν = 100 mV/s; all concentrations given are final concentrations.

The GSH-concentration dependence of voltammetric characteristics for a GCE/CoPc electrode obtained in the presence of 10 µM Hcys is presented in Figure 14. The relationships ipa = f(CGSH) and ipc = f(CGSH) are linear in the lower concentration range (CGSH < 2 mM) and show a decreased slope (sensitivity) ∂ip/∂CGSH at higher GSH concentrations. Voltammograms for GSH solutions containing 10 µM Hcys, recorded for different potential scan rates, are presented in Figure 14b. In the presence of Hcys, the dependence of GSH oxidation peak current still is linear with square root of the scan rate v, indicating that the transport of GSH from solution is involved in the electrode process. The same concerns to GSSG-reduction peak current. The experiments for GSH solutions containing 100 µM Hcys have also been performed. The GSH-concentration dependence of voltammetric characteristics for a GCE/CoPc electrode obtained in the presence of 100 µM Hcys is presented in Figure 15a. The relationships ipa = f(CGSH) and ipc = f(CGSH) are linear in the lower concentration range (CGSH < 2 mM) and show a decreased slope (sensitivity) ∂ip/∂CGSH at higher GSH concentrations. Voltammograms for GSH solutions containing 100 µM Hcys, recorded for different potential scan rates, are presented in Figure 15b. In the presence of Hcys, the dependence of GSH oxidation peak current is still linear with square root of the scan rate v, indicating that the transport of GSH from solution is involved in the electrode process. The same concerns to GSSG-reduction peak current.

344 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 14. (a) Cyclic voltammograms for a GCE/CoPc electrode obtained for GSH solutions containing 10 µM Hcys, CGSH [mM]: (1) 0, (2) 0.4, (3) 0.79, (4) 1.6, (5) 2.3, (6) 3.1, (7) 3.85; (8) 4.5; 50 mM phosphate buffer, pH 7.43; ν = 100 mV/s. INSET: Dependence of i vs. CGSH. (b) Cyclic voltammograms for a GCE/CoPc in 4.5 mM GSH solution containing 10 µM Hcys, obtained for different scan rates v [mV/s]: (1) 25, (2) 50, (3) 75, (4) 100, (5) 200, (6) 300, (7) 400, (8) 500; 50 mM phosphate buffer, pH 7.43. (see color insert)

Figure 15. (a) Cyclic voltammograms for a GCE/CoPc electrode obtained for GSH solutions containing 100 µM Hcys, CGSH [mM]: (1) 0, (2) 0.4, (3) 0.79, (4) 1.6, (5) 2.3, (6) 3.1, (7) 3.85; (8) 4.5; ν = 100 mV/s. (b) Cyclic voltammograms for a GCE/CoPc in 4.5 mM GSH solution containing 100 µM Hcys, obtained for v [mV/s]: (1) 25, (2) 50, (3) 75, (4) 100, (5) 200, (6) 300, (7) 400, (8) 500. Other conditions: 50 mM phosphate buffer, pH 7.43.

345 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Since many oxidation processes of bioorganic compounds proceed faster in alkaline than in neutral solutions, we have tested the CoPc catalyst performance in Britton-Robinson buffer of pH 12. The results are presented in Figure 16 for a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) in GSH solutions containing 10 µM Hcys. The anodic oxidation peak is better defined than that observed at lower pH and the peak potential Epa = +0.06 V is also lower (Epa = +0.20 V for pH = 7.43). The oxidation onset is seen at potentials as low as E = -0.26 V. However, the peak current values relative to the background capacitive current are lower than those at lower pH. As a consequence, the limit of detection is lower for neutral pH. For this reason, the neutral pH has been selected as better suited for the GSH analysis.

Figure 16. Cyclic voltammograms for a cobalt phthalocyanine-modified glassy carbon electrode (GCE/CoPc) obtained for GSH solutions containing 10 µM Hcys, CGSH [mM]: (1) 0, (2) 0.79, (3) 1.6, (4) 2.3, (5) 3.1, (6) 3.85; (7) 4.5; 40 mM B-R buffer, pH 12; ν = 100 mV/s.

The comparison of various calibration plots for GSH, Hcys and Cys is presented in Figure 17. The highest sensitivity is obtained for Hcys, followed by Cys, and the least sensitivity is observed for GSH. This is probably due to the larger size of GSH molecule and also its multiple local positive charges which are repelled by the Co+1/+2 cation in CoPc core being the catalytic center of the electron exchange process. It is also seen that the competition from Hcys in GSH determination vanishes at lower GSH concentrations.

346 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

Figure 17. Dependence of ipa vs. Canalyte for different biomarkers of oxidative stress, obtained for a phosphate buffer pH 7.43, on a CoPC-modiefied GCE electrode, for: (1) Hcys, (2) Cys, (3) GSH, (4) GSH in the presence of 10 µM Hcys, v = 100 mV/s.

Conclusions In the absence of biocatalytic effects, the redox reactivity of GSH/GSSG couple on solid electrodes is strongly hindered. We have shown that the rates of redox reactions of glutathione, as well as other biothiols: cysteine and homocysteine, can be enhanced on electrocatalytic cobalt phthalocyanine (CoPc) monolayer film electrodes, enabling voltammetric detection of these important biomarkers of oxidative stress. We have demonstrated that a strong competition between GSH, Cys, and Hcys exists due to the competitive charge-transfer complex formation. We have previously observed a strong ligand competion in thiol-capped Au nanoparticle assembly processes studied using UV-Vis and resonance elastic light scattering spectroscopies. While the strong competition of thiols prevents straightforward analysis of multicomponent mixtures, the detailed investigations indicate that the analysis can be carried out using the standard addition method taking into account active matrix effects which cannot be neglected.

Acknowledgments This work was partially supported by the U.S. DoD grant No. AS073218.

347 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

References 1. 2.

3.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

4. 5. 6. 7. 8. 9. 10. 11. 12.

13.

14. 15. 16. 17. 18. 19. 20.

21. 22. 23.

Sen, S.; Chakraborty, R.; Sridhar, C.; Reddy, Y. S. R.; De, B. Int. J. Pharm. Sci. Rev. Res. 2010, 3, 91–100. Hepel, M.; Stobiecka, M., Detection of Oxidative Stress Biomarkers Using Functional Gold Nanoparticles. In Fine Particles in Medicine and Pharmacy; Matijevic, E., Ed.; Springer Science Publ.: New York, 2012; pp 241-281. Poljsak, B., Decreasing Oxidative Stress and Retarding the Aging Process; Nova Sci. Publishing: New York, 2010. Tardif, J.-C. Cardiology Rounds 2003, 7, 9. Lakshmi, S. V. V.; Padmija, G.; Kuppusamy, P.; Kutala, V. K. Indian J. Biochem. Biophys. 2009, 46, 421–440. Krishnan, C. V.; Garnett, M.; Chu, B. Int. J. Electrochem. Sci. 2008, 3, 1348–1363. Nikam, S.; Nikam, P.; Ahaley, S. K.; Sontakke, A. V. Indian J. Clin. Biochem. 2009, 24, 98–101. Singh, R. P.; Sharad, S.; Kapur, S. JIACM 2004, 5, 218–25. Perluigi, M.; Butterfield, D. A. Expert Rev. Proteomics 2011, 8, 427–429. Zana, M.; Janka, Z.; Kalman, J. Neurobiol. Aging 2007, 28, 648–676. James, S. J.; Cutler, P.; Melnyk, S.; Jernigan, S.; Janak, L.; Gaylor, D. W.; Neubrander, J. A. Am. J. Clin. Nutr. 2004, 80, 1611–1617. James, S. J.; Melnyk, S.; Jernigan, S.; Cleves, M. A.; Halsted, C. H.; Wong, D. H.; Cutler, P.; Bock, K.; Boris, M.; Bradstreet, J. J.; Baker, S. M.; Gaylor, D. W. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2006, 141B, 947–956. Stobiecka, M.; Prance, A.; Coopersmith, K.; Hepel, M. Antioxidant effectiveness in preventing paraquat-mediated oxidative DNA damage in hte presence of H2O2. In Oxidative Stress: Diagnostics, Prevention and Therapy; Andreescu, S., Hepel, M., Eds.; Oxford University Press: Oxford, 2012; Vol. 1083, pp 211−233. Hepel, M.; Stobiecka, M.; Peachey, J.; Miller, J. Mutat. Res. 2012, 735, 1–11. Moussa, S. A. Romanian J. Biophys. 2008, 18, 225–236. Wiernsperger, N. F. Diabetes Metab. 2003, 29, 579–585. Markesbery, W. R. Free Radical Biol. Med. 1997, 23, 134–147. Perry, G.; Cash, A. D.; Smith, M. A. J. Biomed. Biotechnol. 2002, 2 (3), 120–123. Yves, C. Am. J. Clin. Nutr. 2000, 71, 621S–9S. Hamilton, C. A.; H. Miller, W. H.; Al-Benna, S.; Brosnan, M. J.; Drummond, R. D.; McBride, M. W.; Dominiczak, A. F. Clin. Sci. 2004, 106, 219–234. Madamanchi, N. R.; Vendrov, A.; Runge, M. S. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 29–38. Galle, J. Nephrol. Dial. Transplant. 2001, 16, 2135–2137. Himmelfarb, J.; McMonagle, E.; Freedman, S.; Klenzak, J.; McMenamin, E.; Le, P.; Pupim, L. B.; Ikizler, T. A.; PicardGroup. J. Am. Soc. Nephrol. 2004, 15, 2449–2456. 348 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

24. Halliwell, B. Biochem. J. 2007, 401, 1–11. 25. Tandon, V. R.; Sharma, S.; Mahajan, A.; Bardi, G. H. JK Science 2005, 7, 1–3. 26. Masella, R.; Di Benedetto, R.; Vari, R.; Filesi, C.; Giovannini, C. J. Nutr. Biochem. 2005, 16, 577–586. 27. Kohen, R.; Nyska, A. Toxicol. Pathol. 2002, 30, 620–650. 28. Hepel, M.; Stobiecka, M. Detection of Oxidative Stress Biomarkers Using Novel Nanostructured Biosensors. In New Perspectives in Biosensors Technology and Applications; Serra, P. A., Ed.; INTECH: Vienna, 2011; pp 343−372. 29. Mansoor, M. A.; Svardal, A. M.; Ueland, P. M. Anal. Biochem. 1992, 200, 218–229. 30. Kleinman, W. A.; Richie, J. P. Biochem. Pharmacol. 2000, 60, 19–29. 31. Harfield, J. C.; Batchelor-McAuley, C.; Compton, R. G. Analyst 2012, 137, 2285–2296. 32. White, P. C.; Lawrence, N. S.; Davis, J.; Compton, R. G. Electroanalysis 2002, 14, 89–98. 33. Cereser, C.; Guichard, J.; Drai, J.; Bannier, E.; Garcia, I.; Boget, S.; Parvaz, P.; Revol, A. J. Chromatogr. B 2001, 752, 123–132. 34. Kusmierek, K.; Glowacki, R.; Bald, E. Anal. Bioanal. Chem. 2006, 385, 855–860. 35. Refsum, H.; Ueland, P. D.; Nygård, P.; Vollset, S. E. Annu. Rev. Med. 1998, 49, 31–62. 36. Tcherkas, Y. V.; Denisenko, A. D. J. Chromatogr. A 2001, 913, 309–313. 37. Stabler, S. P.; Marcell, P. D.; Rodell, E. R.; Allen, R. H.; Savage, D. G.; Lindenbaum, J. J. Clin. Invest. 1988, 81, 466–474. 38. Spãtaru, N.; Sarada, B. V.; Popa, E.; Tryk, D. A.; Fujishima, A. Anal. Chem. 2001, 73, 514–519. 39. Stobiecka, M.; Deeb, J.; Hepel, M. Biophys. Chem. 2010, 146, 98–107. 40. Stobiecka, M.; Hepel, M. Sens. Actuators, B 2010, 149, 373–380. 41. Xiao, Q.; Zhang, L.; Lu, C. Sens. Actuators, B 2012, 166-167, 650–657. 42. Wang, J.; Li, Y. F.; Huang, C. Z.; Wua, T. Anal. Chim. Acta 2008, 626, 37–43. 43. Sun, S. K.; Wang, H. F.; Yan, X. P. Chem. Commun. 2011, 47, 3817–3819. 44. Leesutthiphonchai, W.; Dungchai, W.; Siangproh, W.; Ngamrojnavanich, N.; Chailapakul, O. Talanta 2011, 85, 870–876. 45. Hepel, M.; Stobiecka, M. J. Photochem. Photobiol. A 2011, 225, 72–80. 46. Xu, H.; Hepel, M. Anal. Chem. 2011, 83, 813–819. 47. Tan, H.; Chen, Y. J. Biomed. Opt. 2012, 17, 017001. 48. Stobiecka, M.; Molinero, A. A.; Chalupa, A.; Hepel, M. Anal. Chem. 2012, 84, 4970–4978. 49. Stobiecka, M.; Hepel, M. Biosens. Bioelectron. 2011, 26, 3524–3530. 50. Stobiecka, M.; Deeb, J.; Hepel, M. Electrochem. Soc. Trans. 2009, 19, 15–32. 51. Çubukçu, M.; Ertaş, F. N.; Anık, Ü. Microchim. Acta 2013, 180, 93–100. 52. Ndamanisha, J. C.; Bai, J.; Qi, B.; Guo, L. Anal. Biochem. 2009, 386, 79–84. 349 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

53. Pereira-Rodrigues, N.; Cofre, R.; Zagal, J. H.; Bedioui, F. Bioelectrochem. 2007, 70, 147–154. 54. Sehlotho, N.; Griveau, S.; Ruillé, N.; Boujtita, M.; Nyokong, T.; Bedioui, F. Mater. Sci. Eng. C 2008, 28, 606–612. 55. Guttierrez, A. P.; Argote, M. R.; Griveau, S.; Zagal, J. H.; Granados, S. G.; Ordas, A. A.; Bedioui, F. J. Chil. Chem. Soc. 2012, 52, 1244–1247. 56. Raoof, J. B.; Ojani, R.; Baghayeri, M. Sens. Actuators, B 2009, 143, 261–269. 57. Inoue, T.; Kirchhoff, J. R. Anal. Chem. 2000, 72, 5755–5760. 58. Joshi, K. A.; Pandey, P. C.; Chen, W.; Mulchandani, A. Electroanalysis 2004, 16, 1938–1943. 59. Han, H.; Tachikawa, H. Front. Biosci. 2005, 10, 931–939. 60. Lee, P. T.; Compton, R. G. Electroanalysis 2013, 25, 1613–1620. 61. Lee, P. T.; Ward, K. R.; Tschulik, K.; Chapman, G.; Compton, R. G. Electroanalysis 2014, 26, 366–373. 62. Lee, P. T.; Lowinsohn, D.; Compton, R. G. Sensors 2014, 14, 10395–10411. 63. Richie, J. P.; Lang, C. A. Anal. Biochem. 1987, 163, 9–15. 64. Gong, K.; Zhu, X.; Zhao, R.; Xiong, S.; Mao, L.; Chen, C. Anal. Chem. 2005, 77, 8158–8165. 65. Abiman, P.; Wildgoose, G. G.; Compton, R. G. Electroanalysis 2007, 19, 437–444. 66. Arduini, F.; Majorani, C.; Amine, A.; Moscone, D.; Palleschi, G. Electrochim. Acta 2011, 56, 4209–4215. 67. Gong, Z. X.; Li, H. J. Electrochem. Soc. 2000, 147, 238–241. 68. Houze, P.; Gamra, S.; Madelaine, I.; Bousquet, B.; Gourmel, B. J. Clin. Lab. Anal. 2001, 15, 144–153. 69. Carvalho, F. D.; Remião, F.; Valet, P.; Timbrell, J. A.; Bastos, M. I.; Ferreira, M. A. Biomed. Chromatogr. 2005, 8, 134–136. 70. Chen, J.; He, Z.; Liu, H.; Cha, C. J. Electroanal. Chem. 2006, 588, 324–330. 71. Pacsial-Ong, E. J.; McCarley, R. L.; Wang, W.; Strongin, R. M. Anal. Chem. 2006, 78, 7577–7581. 72. Vandeberg, P. J.; Johnson, D. C. Anal. Chem. 1993, 65, 2713–2718. 73. Winters, R. A.; Zukowski, J.; Ercal, N.; Matthews, R. H.; Spitz, D. R. Anal. Biochem. 1995, 227, 14–21. 74. Mazloum-Ardakani, M.; Sheikh-Mohseni, M. A.; Mirjalili, B. F. Electroanalysis 2013, 25, 2021–2029. 75. Salehzadeh, H.; Mokhtari, B.; Nematollahi, D. Electrochim. Acta 2014, 123, 353–361. 76. Nekrassova, O.; White, P. C.; Threlfell, S.; Hignett, G.; Wain, A. J.; Lawrence, N. S.; Davis, J.; Compton, R. G. Analyst 2002, 127, 797–802. 77. Scampicchio, M.; Lawrence, N. S.; Arecchi, A.; Mannino, S. Electroanalysis 2007, 19, 2437–2443. 78. Hignett, G.; Threlfell, S.; Wain, A. J.; Lawrence, N. S.; Wilkins, S. J.; Davis, J.; Compton, R. G.; Cardosi, M. F. Analyst 2001, 126, 353–357. 79. Seymour, E. H.; Wilkins, S. J.; Lawrence, N. S.; Compton, R. G. Anal. Lett. 2002, 35, 1387–1399. 350 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.

Downloaded by PURDUE UNIV on November 2, 2015 | http://pubs.acs.org Publication Date (Web): October 13, 2015 | doi: 10.1021/bk-2015-1200.ch014

80. White, P. C.; Lawrence, N. S.; Tsai, Y. C.; Davis, J.; Compton, R. G. Mikrochim. Acta 2001, 137, 87–91. 81. Liu, X.; Lv, H.; Sun, Q.; Zhong, Y.; Zhao, J.; , J.Fu; Lin, M.; Wang, J. Anal. Lett. 2012, 45, 2246–2256. 82. Lowinsohn, D.; Lee, P. T.; Compton, R. G. Int. J. Electrochem. Sci. 2014, 9, 3458–3472. 83. Wang, W. H.; Escobedo, J. O.; Lawrence, C. M.; Strongin, R. M. J. Am. Chem. Soc. 2004, 126, 3400. 84. Lu, C.; Zu, Y. B. Chem. Commun. 2007, 3871. 85. Chen, H. L.; Zhao, Q.; Wu, Y. B.; Li, F. Y.; Yang, H.; Yi, T.; Huang, C. H. Inorg. Chem. 2007, 46, 11075. 86. Lim, I. I. S.; Ip, W.; Crew, E.; Njoki, P. N.; Mott, D.; Zhong, C. J.; Pan, Y.; Zhou, S. Q. Langmuir 2007, 23, 826. 87. Lim, I. I. S.; Mott, D.; Ip, W.; Njoki, P. N.; Pan, Y.; Zhou, S. Q.; Zhong, C. J. Langmuir 2008, 24, 8857. 88. Stobiecka, M.; Coopersmith, K.; Hepel, M. J. Colloid Interface Sci. 2010, 350, 168–177. 89. Atkins, P. W.; Friedman, R. S. Molecular Quantum Mechanics; Oxford University Press: Oxford, 2004. 90. Hehre, W. J.; Radon, L.; Schleyer, P. R.; Pople, J. A. Ab-initio Molecular Orbital Theory; Wiley: New York, 1985. 91. Santos da Silva, I.; Araújo, M. F. A.; Ferreira, H. A.; de Jesus Gomes Varela, J., Jr.; Tanaka, S. M. C. N.; Tanaka, A. A.; Angnes, L. Talanta 2011, 83, 1701–1706. 92. Luz, R. C. S.; Moreira, A. B.; Damos, F. S.; Tanaka, A. A.; Kubota, L. T. J. Pharm. Biomed. Anal. 2006, 42, 184–191. 93. Schafer, F. Q.; Buettner, G. R. Free Radical Biol. Med. 2001, 30, 1191–1212. 94. Hiroi, M.; Ogihara, T.; Hirano, K.; Hasegawa, M.; Morinobu, T.; Tamai, H.; Niki, E. Free Radical Biol. Med. 2005, 38, 1057–1072.

351 In Oxidative Stress: Diagnostics, Prevention, and Therapy Volume 2; Hepel, et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2015.